Next Article in Journal
Astragaloside IV Attenuates Programmed Death-Ligand 1-Mediated Immunosuppression during Liver Cancer Development via the miR-135b-5p/CNDP1 Axis
Previous Article in Journal
Unveiling the Role of Tumor-Infiltrating T Cells and Immunotherapy in Hepatocellular Carcinoma: A Comprehensive Review
Previous Article in Special Issue
Renal Function Parameters in Distinctive Molecular Subtypes of Prostate Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Castration-Resistant Prostate Cancer: From Uncovered Resistance Mechanisms to Current Treatments

by
Thi Khanh Le
1,2,†,
Quang Hieu Duong
1,2,3,†,
Virginie Baylot
1,2,
Christelle Fargette
2,4,
Michael Baboudjian
1,2,5,
Laurence Colleaux
6,
David Taïeb
1,2,4 and
Palma Rocchi
1,2,*
1
Centre de Recherche en Cancérologie de Marseille—CRCM, Inserm UMR1068, CNRS UMR7258, Aix-Marseille University U105, 13009 Marseille, France
2
European Center for Research in Medical Imaging (CERIMED), Aix-Marseille University, 13005 Marseille, France
3
Vietnam Academy of Science and Technology (VAST), University of Science and Technology of Hanoi (USTH), Hanoi 10000, Vietnam
4
Department of Nuclear Medicine, La Timone University Hospital, Aix-Marseille University, 13005 Marseille, France
5
Department of Urology AP-HM, Aix-Marseille University, 13005 Marseille, France
6
Faculté de Médecine Timone, INSERM, MMG, U1251, Aix-Marseille University, 13385 Marseille, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2023, 15(20), 5047; https://doi.org/10.3390/cancers15205047
Submission received: 28 August 2023 / Revised: 26 September 2023 / Accepted: 12 October 2023 / Published: 19 October 2023
(This article belongs to the Special Issue Prostate Cancer Progression)

Abstract

:

Simple Summary

Castration-resistant prostate cancer (CRPC) remains a significant medical challenge, even with recent advancements in diagnosis and treatment. To improve patient outcomes, it is important to understand the underlying mechanisms of resistance to treatments and develop new therapeutic approaches. This review provides a brief summary of the current knowledge on the mechanisms that contribute to CRPC progression, including both androgen receptor (AR)-dependent and AR-independent pathways. It also discusses approved and currently investigated treatment options to treat patients with CRPC, such as novel chemotherapies, radiation therapy, immunotherapy, PARP inhibitors, and potential combined therapeutic strategies.

Abstract

Prostate cancer (PC) is the second most common cancer in men worldwide. Despite recent advances in diagnosis and treatment, castration-resistant prostate cancer (CRPC) remains a significant medical challenge. Prostate cancer cells can develop mechanisms to resist androgen deprivation therapy, such as AR overexpression, AR mutations, alterations in AR coregulators, increased steroidogenic signaling pathways, outlaw pathways, and bypass pathways. Various treatment options for CRPC exist, including androgen deprivation therapy, chemotherapy, immunotherapy, localized or systemic therapeutic radiation, and PARP inhibitors. However, more research is needed to combat CRPC effectively. Further investigation into the underlying mechanisms of the disease and the development of new therapeutic strategies will be crucial in improving patient outcomes. The present work summarizes the current knowledge regarding the underlying mechanisms that promote CRPC, including both AR-dependent and independent pathways. Additionally, we provide an overview of the currently approved therapeutic options for CRPC, with special emphasis on chemotherapy, radiation therapy, immunotherapy, PARP inhibitors, and potential combination strategies.

1. Introduction

Prostate cancer (PC) is the second most prevalent cancer in men worldwide, representing 15% of all cancers diagnosed in males [1]. In the US, it is the most frequently diagnosed cancer and the second leading cause of cancer deaths in men. An estimated 288,300 new PC cases and 34,700 deaths from the disease are predicted to occur in the US in 2023 [2]. Patients who have advanced PC initially respond very well to androgen deprivation therapy. However, the treatment finally selects cancer cells that relapse into androgen deprivation, leading to the emergence of castration-resistant prostate cancer (CRPC). Patients with metastatic CRPC (mCRPC) have reduced life expectancy, with a median overall survival rate of below 2 years. The current definition CRPC, according to the European Association of Urology (EAU), is based on biochemical and/or clinical parameters demonstrating cancer progression in a castration environment (testosterone level < 50 ng/dL or 1.7 nmol/L) and standards proposed by Prostate Cancer Working Group 3 (PCWG3) [3] and/or the Response Evaluation Criteria in Solid Tumors (RECIST 1.1) [4]. Accordingly, biological progression refers to a condition where there are three consecutive increases in PSA (Prostate Specific Antigen), each at least one week apart, resulting in two increases of at least 50% over the lowest point, and a PSA level greater than 2 ng/mL. Radiographic progression is defined by the appearance of at least two new lesions on a bone scan or the progression of a measurable lesion according to the RECIST [5].
Over the past decade, there has been significant improvement in the treatment options for CRPC. Several drugs that increase overall patient survival, including androgen-receptor (AR) signaling inhibitors (such as abiraterone acetate, enzalutamide, apalutamide, and darolutamide), and radiopharmaceutical therapies (like radium-223 and 177Lu-PSMA-617) have been approved [6]. Advancements in precision medicine have led to the discovery of distinct subtypes of prostate cancer and the identification of genetic changes that predict the effectiveness of specific treatments [7,8]. In this review, we will provide an overview of the current understanding of the mechanisms that lead to CRPC, including both AR-dependent and -independent pathways. This review also summarizes the therapeutic options currently available for treating CRPC.

2. Mechanisms Underlying Castration-Resistant Prostate Cancer

There are currently many treatment options available for CRPC patients; however, the responses to these therapies are still limited. Knowledge of the underlying mechanisms associated with the CRPC phenotype can help to identify new promising therapeutics.

2.1. Androgen Receptor Overexpression

Androgen Receptor (AR) overexpression enables the survival and proliferation of tumor cells in limited-androgen conditions during androgen suppression treatment. Elevated AR expression at both the mRNA and protein levels has been frequently observed in CRPC. Amplification of the AR has been observed in 70% of cases of CRPC and is associated with significantly increased AR mRNA expression [9]. David A. et al. reported the existence of an enhancer that is amplified in 81% of castration-resistant metastatic patients. This enhancer has the potential to increase the expression of the AR gene independently of AR locus amplification in response to first-line androgen deprivation therapy (ADT) [9]. It should be noted that AR amplification is only seen in prostate tumors that have been exposed to androgen deprivation, indicating that AR amplification is a consequence of hormone therapy [10]. Additionally, epigenetic modifications, such as DNA methylation, histone acetylation, and miRNA modulation, could lead to AR overexpression in CRPC (Figure 1) [11].

2.2. AR Mutations

AR gene mutations have been found in around 10–20% of CRPC cases, while they are rarely observed in localized PC (0–4%) [12,13]. Most recurrent AR mutations are found in the ligand-binding domain (LBD) and/or cofactor binding regions [14]. These alterations could decrease the specificity of the AR for its main ligand, androgen; however, they allow AR specificity to extend to other hormones, such as progesterone and estrogen [12,15,16,17,18]. In addition, they could also disrupt the effectiveness of AR antagonists, causing them to act as AR agonists and resulting in resistance to treatments [19]. It has been reported that the H875Y mutation in CWR22 causes cells to exhibit altered ligand specificity. This mutant AR has shown transcriptional activity in response to testicular androgens, similarly to the wild-type AR. However, it differs from the wild type in its additional activation by adrenal androgens, dehydroepiandrosterone, and the antiandrogen hydroxyflutamide, as well as by estradiol and progesterone [20]. Along with AR coding mutations, several AR splicing mutations, leading to the abnormal expression of splice variants such as AR3 (also called AR-V7), AR4, and AR5, which lack the ligand-binding domain, are frequently observed in CRPC [21]. These variants encode a shorter receptor that lacks the ligand-binding domain and thereby constitutively activates AR pathways [22].

2.3. Alteration in AR CoRegulators

As a transcription factor (TF), the AR cooperates with a variety of regulatory proteins to form a productive transcription complex. These coregulators can either enhance (coactivators) or depress (cosuppressors) the transcriptional activity of the AR, thereby modulating the expression of androgen-regulated genes. As a consequence, alteration in the expression of AR coregulators can give a survival advantage to the cancer cells during castration therapy, promoting CRPC progression [23]. Several well-documented AR coactivators, including SRC1, SRC2, SRC3, ARA70, PIAS1, and Tip60, interact with the AR and enhance its transcriptional activity [24]. Levels of TIF2 and SRC1 expression have been observed to rise alongside the expression of the AR during the progression of prostate cancer cells following androgen deprivation. A previous study has also shown that these proteins are overexpressed in the majority of CRPC samples. The overexpression of TIF2 enhances the transcriptional activation of the AR in response to weaker steroids, such as adrenal androgens (DHEA and androstenedione), in two cell lines carrying AR mutations: LNCaP (T877A) and CWR22 (H874Y) [25]. The steroid receptor coactivator-3 (SRC3), also called amplified-in-breast cancer-1 (AIB1), has been identified as a potent coactivator of the hormone-activated AR [26] and has been demonstrated to promote CRPC progression [27]. Conversely, two well-known corepressors of the AR, NCoR (nuclear receptor corepressor) and SMRT (silencing mediator for retinoid and thyroid hormone receptors), have been shown to interact with the AR and decrease its transcriptional activity triggered by dihydrotestosterone [28,29,30].

2.4. Increased Steroidogenic Signaling Pathways

Accumulating evidence has reported that PC cells could survive after castration therapy by regulating intracrine androgen synthesis within the prostate. A typical example is the 5α-reductase enzyme, which catalyzes the conversion of testosterone (the most abundant free androgen) to the higher-affinity ligand 5α-dihydrotestosterone (DHT). Upon ADT, intraprostatic testosterone and DHT levels do not decline as markedly as do serum levels after ADT, buffering tumor cells from the loss of testicular androgen [31]. Increased 5α-reductase enzyme levels result from a polymorphism where there is a substitution of a valine with a leucine at the codon 89 [32]. Intraprostatic androgens can also be synthesized from cholesterol or other molecular precursors, such as DHEA (dehydroepiandrosterone). Indeed, DHEA could be converted into androstenedione, a substrate for conversion to testosterone. Another report has shown that CRPC samples have higher levels of expression of a variety of enzymes involved in de novo steroid synthesis, such as FASN, CYP17A1, CYP19A1, and UGT2B17, compared to primary PC [33].

2.5. Outlaw Pathways

In addition of being activated primarily by endogenous androgen ligands, the AR protein can also be stimulated through ligand-independent mechanisms, which are known as outlaw pathways. CRPC can thus be induced by the interactions between the cytosolic AR and many molecules such as different growth factors, cytokines, and kinases. Several growth factors, such as insulin-like growth factor-I (IGF-I), keratinocyte growth factor (KGF) and the epidermal growth factor (EGF), enable the stimulation of AR transcription activity at limited androgen levels or even in the absence of androgens [17]. Particularly, IGF1 has been reported to interact with the AR, facilitate AR translocation, and modulate the transcription of androgen-responsive genes. Moreover, it has been demonstrated to upregulate the expression of the AR coactivator TIF2 [34]. Along with growth factors, many cytokines, such as IL-6 and IL-8, have proven their ability to activate AR signaling in a ligand-independent manner [35]. Additionally, receptor tyrosine kinases (RTK) such as HER2/ERBB2 have been shown to be overexpressed in CRPC and restore AR signaling in low-androgen-concentration conditions [25].

2.6. Non-AR-Related Pathways: Bypass Pathways

While the mechanisms described above rely on AR transactivation to promote CRPC progression, alternative survival and growth pathways that are independent of AR activation can also drive androgen deprivation resistance. Of note, many outlaw pathways can also drive CRPC progression in an AR-signaling independent manner. High levels of serum insulin-like growth factor-I (IGF-I) and its receptor (IGF-1R) have been increasingly recognized to play a key role in prostate cancer progression. Binding of IGF-I with IGF-IR enables the stimulation of the transduction of a signal cascade, which, in turn, activates the expression of genes responsible for cellular growth and proliferation [10,23]. The upregulation of Akt pathway has been reported in a variety of human malignancies, including prostate cancer [36]. Indeed, AKT plays a central role in antiapoptotic pathways by phosphorylating and depressing Bcl2-associated agonist of cell death (BAD), a proapoptotic member of the BCL-2 protein family, and pro-caspase 9 [37]. AKT also promotes cellular proliferation by decreasing the expression of p27, a cell cycle inhibitor [38]. Several reports have shown that loss of PTEN, a tumor suppressor that inhibits PI3K/AKT signaling, frequently occurs in CRPC samples [39].
In addition, the heat-shock protein Hsp27 has previously been demonstrated to drive therapy resistance in PC, and its inhibitors, such as Hsp27 ASO (OGX-427) and small interference RNA (siRNA), have shown the ability to enhance chemotherapy. More recently, we have shown that Hsp27 plays a role in promoting the development of CRPC through its chaperone activity by selectively protecting some partner proteins involved in castration resistance, such as eIF4E, TCTP, Menin, and DDX5, from their ubiquitin–proteasome degradation [40,41,42,43]. CRPC has the potential to develop resistance to the antiandrogen enzalutamide by bypassing the AR blockade, which occurs due to increased activity of the glucocorticoid receptor (GR) [44]. Recent studies have indicated that heightened activity of the GR has been linked to the emergence of resistance to antiandrogen therapies in different clinical and preclinical models [44,45]. These findings could possibly be explained by the fact that GR expression is downregulated by AR signaling [46]. Additionally, the GR interacts with and mediates a subset of AR targets [44].

2.7. Evolution of t-NEPC from CRPC

During the initial diagnosis of PC, neuroendocrine prostate cancer (NEPC) is rare and represents only 0.5–2% of all PC cases. However, a larger proportion of treatment-induced neuroendocrine prostate cancer (t-NEPC) has been reported. Recent studies have indicated that approximately 17–30% of CRPC patients, following ADT and other treatments, may experience progression to t-NEPC [47,48]. The majority of evidence suggests that the origin of tNEPC is the transdifferentiation of adenocarcinoma cells into NEPC cells in response to different therapies, including ADT [47,48,49]. The transition to the NEPC phenotype following ADT treatments has been demonstrated in various preclinical models, including cell lines, genetically engineered mouse (GEM), and patient-derived xenografts [50,51,52]. In addition to ADT, other therapies, such as radiation and chemotherapy, have been reported to stimulate the NEPC phenotype [53,54,55]. Moreover, genetic and epigenetic alterations and tumor microenvironments (TMEs) containing a variety of cells and factors could support NEPC transition in CRPC patients [56,57,58,59]. Due to the absence of AR and PSA expression (AR-/PSA-), t-NEPC patients have shown limited responses to existing treatment options. To date, several therapeutics have been proposed for patients with t-NEPC, including cisplatin or carboplatin combined with etoposide or docetaxel.

3. Treatment Options for CRPC

Treatment options depend on several key factors, such as age, health condition, disease stage and grade, responses to previous therapies, and the potential advantages and side effects of the treatment. Besides ADT, other standard therapeutics for CRPC include chemotherapy, immunotherapy, radiation therapy, and PARP inhibitors (Figure 2).

3.1. Androgen Deprivation Therapy (ADT)

Androgens play a key role in regulating the proliferation and survival of prostatic cells. Androgen deprivation therapy aims to decrease the androgen levels produced by the testicular area, thereby preventing them from fueling prostate cancer cells, making the cancer cells shrink or grow slowly. ADT can be achieved either with surgery (pulpectomy) or pharmacologically.
For medical castration, a variety of compounds can be used, such as LHRH agonists and LHRH antagonists. Luteinizing hormone-releasing hormone (LHRH) agonists, which are also named LHRH analogs or gonadotropin-releasing hormone (GnRH) agonists, including leuprolide (Lupron®, Eligard®), goserelin (Zoladex®), triptorelin (Trelstar®), and histrelin (Vantas®). An approved LHRH antagonist for PC treatment is degarelix (Firmagon®) [60]. The luteinizing hormone (LH), which is made by the pituitary gland, transmits a key signal for testosterone production in the testes. Another oral ADT drug, relugolix (ORGOVYX™), has been approved by the FDA for patients with advanced PC [61]. A phase III clinical trial of relugolix (NCT03085095), in men with advanced PC with a 54% lower risk of serious adverse cardiovascular events, showed that it suppressed testosterone levels more quickly and effectively than leuprolide [62]. These drugs enable decreased LH production, thereby lowering the gonadal production of testosterone [63].
Currently, it is highly recommended to combine ADT with androgen synthesis inhibitors (e.g., abiraterone acetat) or with antiandrogen therapies (e.g., enzalutamide, apalutamide, darolutamide) as well as in combination with chemotherapies (docetaxel) [64]. The guidance from the European Association of Urology (EAU), the American Urological Association (AUA), and the National Comprehensive Cancer Network (NCCN) indicates that ADT should be maintained for CRPC patients for several reasons [65,66,67]. In fact, even serum testosterone levels are suppressed significantly by ADTs; this hormone still remains at low levels in the serum, thereby maintaining androgen functions in the prostate and in the prostate tumor microenvironment [68]. Moreover, the maintenance of intratumor androgens and of androgen-responsive genes including the AR and PSA has been previously reported [33,69]. Previous sections mentioned adaptive mechanisms of androgen signaling that contribute to CRPC, including overexpression, mutation, and splice variations of the AR, as well as changes in androgen synthesis pathways. The presence of androgens or AR ligands enables these mechanisms to persist or even intensify, which hinders the effectiveness of other treatments [70].

3.2. Chemotherapy

Different chemotherapeutic compounds have been approved for PC treatment, including taxane (docetaxel (Taxotere®) and cabazitaxel (Jevtana®), mitoxantrone (Novantrone®), and estramustine (Emcyt®). In addition, several other chemotherapy drugs, such as cisplatin, oxaliplatin, and carboplatin, are currently in clinical evaluation for their utilization in PC [71]. Both docetaxel and cabazitaxel are drugs that stabilize microtubules, leading to mitotic arrest and apoptotic cell death. In addition, these drugs act on AR nuclear translocation that depends on intact and dynamic microtubules [72]. Docetaxel and cabazitaxel are delivered intravenously once every three weeks for around 10 cycles [73]. In most cases, docetaxel is used as the first-line chemo treatment and cabazitaxel is considered as a second-line treatment to overcome docetaxel resistance [74].
Another chemotherapy drug, mitoxantrone, was approved in 1996 by the FDA to treat CRPC patients. In a phase III trial, 161 patients with symptomatic mCRPC were randomly assigned to receive mitoxantrone plus prednisone versus prednisone alone. The primary study outcome, pain relief response, was achieved in 29% of the patients treated with mitoxantrone versus 12% of the patients treated with prednisone, with no survival benefit observed [75]. Mitoxantrone could cause DNA damage by generating DNA breaks through its binding-to-DNA sequence. It could also act as a topoisomerase II inhibitor, consequently disrupting the DNA replication process [76].
Chemotherapies target rapidly dividing cells to fight cancer, but other constantly fast-dividing cells, like those in the bone marrow, mouth, and intestines, as well as hair follicles, can also be affected, causing important side effects. The type, dose, and duration of chemotherapy determine the undesirable effects. Common ones include hair loss, mouth sores, loss of appetite, nausea, vomiting, diarrhea, increased infection risk, easy bruising or bleeding, and fatigue. These side effects typically disappear after treatment, and some drugs can help to manage them [77].

3.3. Immunotherapy

The purpose of immunotherapies is to boost patient’s own immune system to combat cancer cells. To date, three immunotherapies have been approved by the FDA for the treatment of patients with CRPC, including an immune-cell-based vaccine and two immune checkpoints inhibitors (ICI) [78]. Prostate cancer cells express several proteins, including PSA and prostatic acid phosphatase (PAP), that have been identified as potential targets for antigen-based vaccines for advanced prostate cancer [79]. Sipuleucel-T (first approved 29 April 2010), the first FDA-approved cellular immunotherapy, is produced by collecting dendritic cells from patients. These cells are then activated by incubating them ex vivo with PA2024, a recombinant PAP target fused with the granulocyte–macrophage colony-stimulating factor (GM-CSF). Finally, the activated dendritic cells are reinjected into patients, serving as an autologous cellular vaccine that elicits T-cell immune responses targeting the PAP antigen [80] (Figure 2B). Sipuleucel-T has been tested in several phase III clinical trials (NCT00065442, NCT00005947, and NCT01133704) and received US FDA approval (Provenge®; Dendreon, Inc.; Seattle, WA, USA) [81,82]. In the IMPACT trial (NCT00065442), the median survival of patients treated with Sipuleucel-T was 25.8 months, compared with 21.7 months for patients treated with the placebo, and no difference in the time to objective disease progression was observed [83].
The two other FDA-approved immunotherapies for advanced solid tumors including CRPC are pembrolizumab (first approved 4 September 2014) and dostarlimab (first approved 22 April 2021). They are both ICIs in the form of monoclonal antibodies (mAbs). These drugs inhibit the programmed death 1 pathway (PD-1) by blocking the interaction between the PD-1 receptor and its ligands (PD-L1 and PD-L2), which in turn activates the antitumor immune response [84]. Pembrolizumab and dostarlimab have been approved for patients with unresectable or metastatic microsatellite instability-high or mismatch repair-deficient (dMMR) tumors [78,85]. A CRPC patient who was resistant to abiraterone, docetaxel, and cabazitaxel reportedly experienced a significant decrease in PSA levels, as well as reductions in tumor size and the metastatic lymph nodes, after undergoing pembrolizumab therapy [86]. Furthermore, a phase II single-arm trial (NCT02312557) found that pembrolizumab in combination with enzalutamide provides deep and durable responses for mCRPC patients, regardless of tumor PD-L1 expression or DNA repair defects [87].
Immunotherapies like sipuleucel-T and ICIs offer an appealing alternative to CRPC standard treatments such as chemo- and hormone therapy. However, the benefit of immunotherapies in patients with CRPC remains modest. These observations may be the result of the complexity and heterogeneity of the PC microenvironment. Clearly, a better understanding of the mechanisms underlying immune escape in PC is needed to find which immunotherapeutic targets could be beneficial for the patients and identify the optimal timing to introduce these compounds [88]. This includes identifying and targeting specific antigens on prostate cancer cells, as well as finding ways to overcome the immunosuppressive tumor microenvironment. Additionally, optimal combinations between immunotherapies and standard anti-cancer drugs to treat CRPC are still under investigation.

3.4. Radiation Therapy

3.4.1. PSMA-Targeting Radioligand Therapy

The most efficient and recent strategy for CRPC treatment relies on the use of a prostate-specific membrane antigen (PSMA)-targeting radiolabeled molecule. The PSMA is moderately expressed in several normal tissues. However, its expression is considerably elevated in prostate cancer, regardless of the stage of the disease [89]. As 90% of CRPC cases overexpress the PSMA, they can be treated with PSMA radioligand therapy (RLT). In addition, PSMA also has pro-proliferative functions as the cleavage of glutamate by the PSMA activates the PI3K/AKT/mTOR pathways. High expressions of PSMA are associated with poor clinical outcomes and may be involved in the mechanism by which PC tumors acquire resistance [90,91].
The principle of PSMA RLT is to selectively deliver radiation to cancer cells through the systemic administration of PSMA radioligands such as PSMA-617 [92]. The PSMA RLT scheme relies on regular cycles (every 6 weeks) of fixed activity (7.4 GBq/cycle), similarly to chemotherapy. Dosimetry is not often needed, but it can be mandatory for high-dose therapies. The first step is to select patients that are likely to benefit from this treatment. Patients should be positive for 68Ga- or 18F-PMSA PET (positron emission tomography) scans, both of which act as companion diagnostic tools. In the TheraP trials (NCT03392428), approximately 28% of patients were excluded due to various criteria. These criteria included imaging discrepancies between the results of the 18F-FDG PET and the PSMA PET, as well as low/absent tumor uptake [93]. Lutetium-177 PSMA-617 has been approved by the FDA and EMA for CRPC treatment in patients that have progressed under ASI and one line of taxane-based chemotherapy, based on the excellent results of the VISION trial (NCT03511664). It is expected that promising results before chemotherapy will promote this therapy even earlier, taking into account that such patients will benefit from an earlier triplet that includes chemotherapy [94,95,96,97]. There are many ongoing studies using PSMA RLT, such as PSMAfore (NCT04689828); SPLASH (NCT04647526); and ECLIPSE (NCT05204927). Several mechanisms of resistance to PSMA RLT have been described, such as heterogeneity of dose distribution not covered by the cross-fire effect of 177Lu, visceral organ involvement that is less controlled by radiation, an insufficient cytotoxic effect of 177Lu, and a tumor mutational burden that confers a resistance to radiation damages [98].
Various areas of improvement for 177Lu-PSMA RLT are currently being evaluated: (1) pharmacological manipulation of targets (also called phenotype adjustment) with AR signaling inhibitors (e.g., enzalutamide) [99], (2) validation of image-derived biomarkers for the prediction of responses and prognosis [100,101,102], and (3) enhancement of radiation-induced cell death by combination strategies with immune checkpoint inhibitors or PARP inhibitors [103]. To achieve the intended goal, radiation-induced mutations should involve coding DNA regions in order to induce protein misfolding and the subsequent immunogenic response, (4) replacing 177Lu with alpha emitters such as actinium-225 (NCT04597411) [104,104]. 225Ac-PSMA-617 has been shown to be a viable treatment option with PSA responses in the majority of cases, a finding associated with increased overall survival (OS). It efficacy is, however, decreased in heavily pretreated patients. Substantial bone marrow (one-third) and salivary gland (almost constant) toxicity profiles have been reported but remain tolerable. We urgently need prospective comparative data on actinium–PSMA activity and safety. It is nowadays impossible to establish Ac-225-PSMA as a preferred salvage treatment option for disease progression following 177Lu-PSMA [105]. A significant limitation of PSMA-targeted therapy is the absence of consistent PSMA expression across different metastatic tissues in a patient, especially in the liver, sometimes related to an NE (neuroendocrine) differentiation [106,107,108]. Furthermore, during prostate cancer progression, many patients undergo declines in PSMA expression, making them ineligible for the initiation or continuation of PSMA-directed therapies [97].

3.4.2. Radium-223 Dichloride

Radium-223 dichloride (223Ra; Xofigo®) received the FDA approval for the treatment of bone pain in patients with mCRPC based on the 2013 ALSYMPCA study. When radium-223 atoms decay, they emit four alpha particles that are high-energy and capable of damaging DNA strands, thereby killing cancer cells (Figure 2C). Radium-223 mimics calcium and forms complexes with hydroxyapatite, a mineral found in areas of high bone turnover. This allows it to specifically target bone metastasis [109,110].
The alpha particles emitted by Radium-223 have a short range and a high linear energy transfer. This allows them to kill a small area of tissue intensely (tumors) while sparing most of the normal bone tissue. This is a significant advantage over other radiation therapy options, as it represents a targeted therapeutic approach that minimizes damages in healthy cells. Radium-223 has a dual effect of reducing pathological bone turnover and irradiating tumors. The reduction in pathological bone turnover is especially important for patients with mCRPC, as this condition often leads to the development of bone metastasis, which can cause significant pain and other complications. Radium-223 has been demonstrated to prolong OS in symptomatic patients who have suffered from multiple-bone-metastatic CRPC, without visceral or nodal involvement [111]. CRPC patients with bone metastasis and treated with 223Ra demonstrated significant improved OS in phase III clinical trials [112].

3.5. PARP Inhibitors

The reason for using PARP inhibitors (PARPis) is primarily the high occurrence of genetic mutations in prostate cancer and the concept of synthetic lethality, where the combination of two deficiencies leads to cell death [113]. PARP enzymes participate to DNA repair pathways, specifically in the repair of single-strand breaks via base excision repair, in both normal and cancer cells. The inhibition of PARP enzymes stimulates the conversion of single-strand DNA breaks (SSBs) into double-strand DNA breaks (DSBs) during DNA replication. In normal cells, various DNA damage response (DDR) pathways are activated to repair DSBs, including those of non-homologous end joining (NHEJ), microhomology-mediated end joining (MMEJ), and, most commonly, homologous recombination repair (HRR) [114]. However, in cancer cells with HRR deficiency (HRD), treatment using a PARP inhibitor will ultimately cause buildup of DSBs and result in cell death (Figure 2). Additionally, a PARPi could exert its anticancer activity through its ability to trap PARP1 and PARP2 on damaged DNA sites, which would eventually hinder DNA repair, replication, and transcription [6]. Furthermore, a recent study analyzed the genomic landscape of mCRPC in 197 patients using whole-genome sequencing (WGS). This analysis identified eight genomic subgroups in mCRPC, with 11.2% (22/197) of patients showing significant features of homologous recombination deficiency (HRD), mainly due to biallelic BRCA2 inactivation. Within this subgroup, seven out of twenty-two did not have biallelic BRCA2 inactivation, but four of them had at least one (deleterious) aberration in other BRCAness-related genes [8].
Four PARPis (olaparib, rucaparib, niraparib, and talazoparib) have shown significant antitumor activity against mCRPC. Currently, the FDA has approved two PARPis (olaparib and rucaparib) as monotherapies for metastatic castration-resistant prostate cancer patients [115]. Based on the PROfound study findings, olaparib was approved in 2020 for treatment of mCRPC patients who carry alterations in the genes involved in homologous recombination repair and have experienced progression after androgen receptor signaling inhibitor (ARSI) therapy [116]. Recently, a clinical trial concluded that mCRPC patients with BRCA mutations and who were treated with rucaparib could significantly prolong their progression-free survival time compared to the patients who received the control medication [117]. Furthermore, in May 2023, the FDA approved the first PARPi-based combination for mCRPC patients based on the results of the PROpel trial (NCT03732820). This treatment involves using olaparib, abiraterone, and prednisone as the initial treatment for BRCA-mutated mCRPC patients [118]. Niraparib and talazoparib have only been evaluated as monotherapies in mCRPC patients with DNA repair alterations in phase II trials [119,120]. However, combinations based on these drugs have been authorized for mCRPC patients. The TALAPRO-2 study and the MAGNITUDE trial showed positive results for the combination therapies of talazoparib with enzalutamide and niraparib with abiraterone plus prednisone, respectively, leading to their approval [121,122]. There are currently a number of clinical trials that currently examine the effectiveness of novel PARP inhibitors either as monotherapies or in combination with other agents. These trials, in addition to the PARP inhibitors that have already been approved, offer encouraging prospects for new treatment options for patients with metastatic prostate cancer [123].

4. Conclusions

Although extensive research has been conducted on the mechanisms driving therapy resistance in CRPC, treating this disease remains a significant medical challenge. Therefore, it is crucial to continue studying this theme to identify more effective treatments. Currently, various strategies are used to treat CRPC patients, including maintaining ADT, taxane-based chemotherapy, therapeutic nuclear medicine, the selective use of external beam radiotherapy, PARP inhibitors, and immunotherapy. During treatment, preserving patients’ quality of life (QoL) and bone marrow reserves is of utmost importance. Combining therapies could open up new perspectives.

Author Contributions

Writing—original manuscript preparation: T.K.L., Q.H.D. and V.B.; writing—review and editing: M.B., L.C., C.F., P.R. and D.T.; figure preparation: T.K.L. and Q.H.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by INSERM and grants from the La Ligue (R22030AA), Association pour la Recherche sur les Tumeurs de la Prostate (ARTP) and La Fondation A*Midex -2021 AAP Interdisciplinarité. T.K.L.’s postdoc fellowship was funded by ITMO Cancer (C21033AS). Q.H.D.’s doctoral fellowship was funded by the France Excellence scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferlay, J.; Soerjomataram, I.; Dikshit, R.; Eser, S.; Mathers, C.; Rebelo, M.; Parkin, D.M.; Forman, D.; Bray, F. Cancer Incidence and Mortality Worldwide: Sources, Methods and Major Patterns in GLOBOCAN 2012. Int. J. Cancer 2015, 136, E359–E386. [Google Scholar] [CrossRef]
  2. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer Statistics, 2023. CA. Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef]
  3. Scher, H.I.; Morris, M.J.; Stadler, W.M.; Higano, C.S.; Halabi, S.; Smith, M.R.; Basch, E.M.; Fizazi, K.; Ryan, C.J.; Antonarakis, E.S.; et al. The Prostate Cancer Working Group 3 (PCWG3) Consensus for Trials in Castration-Resistant Prostate Cancer (CRPC). J. Clin. Oncol. 2015, 33, 5000. [Google Scholar] [CrossRef]
  4. Castration-Resistant Prostate Cancer: Mechanisms, Targets and Treatment. SpringerLink. Available online: https://link.springer.com/chapter/10.1007/978-3-319-99286-0_7 (accessed on 24 May 2023).
  5. Eisenhauer, E.A.; Therasse, P.; Bogaerts, J.; Schwartz, L.H.; Sargent, D.; Ford, R.; Dancey, J.; Arbuck, S.; Gwyther, S.; Mooney, M.; et al. New Response Evaluation Criteria in Solid Tumours: Revised RECIST Guideline (Version 1.1). Eur. J. Cancer 2009, 45, 228–247. [Google Scholar] [CrossRef] [PubMed]
  6. Iannantuono, G.M.; Chandran, E.; Floudas, C.S.; Choo-Wosoba, H.; Butera, G.; Roselli, M.; Gulley, J.L.; Karzai, F. Efficacy and Safety of PARP Inhibitors in Metastatic Castration-Resistant Prostate Cancer: A Systematic Review and Meta-Analysis of Clinical Trials. Cancer Treat. Rev. 2023, 120, 102623. [Google Scholar] [CrossRef] [PubMed]
  7. Abida, W.; Armenia, J.; Gopalan, A.; Brennan, R.; Walsh, M.; Barron, D.; Danila, D.; Rathkopf, D.; Morris, M.; Slovin, S.; et al. Prospective Genomic Profiling of Prostate Cancer Across Disease States Reveals Germline and Somatic Alterations That May Affect Clinical Decision Making. JCO Precis. Oncol. 2017, 1, 1–16. [Google Scholar] [CrossRef] [PubMed]
  8. Van Dessel, L.F.; van Riet, J.; Smits, M.; Zhu, Y.; Hamberg, P.; van der Heijden, M.S.; Bergman, A.M.; van Oort, I.M.; de Wit, R.; Voest, E.E.; et al. The Genomic Landscape of Metastatic Castration-Resistant Prostate Cancers Reveals Multiple Distinct Genotypes with Potential Clinical Impact. Nat. Commun. 2019, 10, 5251. [Google Scholar] [CrossRef] [PubMed]
  9. Quigley, D.A.; Dang, H.X.; Zhao, S.G.; Lloyd, P.; Aggarwal, R.; Alumkal, J.J.; Foye, A.; Kothari, V.; Perry, M.D.; Bailey, A.M.; et al. Genomic Hallmarks and Structural Variation in Metastatic Prostate Cancer. Cell 2018, 175, 889. [Google Scholar] [CrossRef]
  10. Saraon, P.; Jarvi, K.; Diamandis, E.P. Molecular Alterations during Progression of Prostate Cancer to Androgen Independence. Clin. Chem. 2011, 57, 1366–1375. [Google Scholar] [CrossRef]
  11. Powell, S.M.; Christiaens, V.; Voulgaraki, D.; Waxman, J.; Claessens, F.; Bevan, C.L. Mechanisms of Androgen Receptor Signalling via Steroid Receptor Coactivator-1 in Prostate. Endocr. Relat. Cancer 2004, 11, 117–130. [Google Scholar] [CrossRef]
  12. Taplin, M.-E.; Bubley, G.J.; Shuster, T.D.; Frantz, M.E.; Spooner, A.E.; Ogata, G.K.; Keer, H.N.; Balk, S.P. Mutation of the Androgen-Receptor Gene in Metastatic Androgen-Independent Prostate Cancer. N. Engl. J. Med. 1995, 332, 1393–1398. [Google Scholar] [CrossRef] [PubMed]
  13. Newmark, J.R.; Hardy, D.O.; Tonb, D.C.; Carter, B.S.; Epstein, J.I.; Isaacs, W.B.; Brown, T.R.; Barrack, E.R. Androgen Receptor Gene Mutations in Human Prostate Cancer. Proc. Natl. Acad. Sci. USA 1992, 89, 6319–6323. [Google Scholar] [CrossRef] [PubMed]
  14. Wyatt, A.W.; Gleave, M.E. Targeting the Adaptive Molecular Landscape of Castration-Resistant Prostate Cancer. EMBO Mol. Med. 2015, 7, 878–894. [Google Scholar] [CrossRef] [PubMed]
  15. Wilding, G.; Chen, M.; Gelmann, E.P. Aberrant Response in Vitro of Hormone-Responsive Prostate Cancer Cells to Antiandrogens. Prostate 1989, 14, 103–115. [Google Scholar] [CrossRef] [PubMed]
  16. Taplin, M.-E.; Bubley, G.J.; Ko, Y.-J.; Small, E.J.; Upton, M.; Rajeshkumar, B.; Balk, S.P. Selection for Androgen Receptor Mutations in Prostate Cancers Treated with Androgen Antagonist1. Cancer Res. 1999, 59, 2511–2515. [Google Scholar]
  17. Culig, Z.; Hobisch, A.; Cronauer, M.V.; Radmayr, C.; Trapman, J.; Hittmair, A.; Bartsch, G.; Klocker, H. Androgen Receptor Activation in Prostatic Tumor Cell Lines by Insulin-Like Growth Factor-1,Keratinocyte Growth Factor and Epidermal Growth Factor. Eur. Urol. 2017, 27, 45–47. [Google Scholar] [CrossRef] [PubMed]
  18. Hara, T.; Miyazaki, J.; Araki, H.; Yamaoka, M.; Kanzaki, N.; Kusaka, M.; Miyamoto, M. Novel Mutations of Androgen Receptor: A Possible Mechanism of Bicalutamide Withdrawal Syndrome. Cancer Res. 2003, 63, 149–153. [Google Scholar] [PubMed]
  19. Watson, P.A.; Arora, V.K.; Sawyers, C.L. Emerging Mechanisms of Resistance to Androgen Receptor Inhibitors in Prostate Cancer. Nat. Rev. Cancer 2015, 15, 701–711. [Google Scholar] [CrossRef]
  20. Tan, J.; Sharief, Y.; Hamil, K.G.; Gregory, C.W.; Zang, D.Y.; Sar, M.; Gumerlock, P.H.; White, R.W.D.; Pretlow, T.G.; Harris, S.E.; et al. Dehydroepiandrosterone Activates Mutant Androgen Receptors Expressed in the Androgen-Dependent Human Prostate Cancer Xenograft CWR22 and LNCaP Cells. Mol. Endocrinol. 1997, 11, 450–459. [Google Scholar] [CrossRef]
  21. Guo, Z.; Yang, X.; Sun, F.; Jiang, R.; Linn, D.E.; Chen, H.; Chen, H.; Kong, X.; Melamed, J.; Tepper, C.G.; et al. A Novel Androgen Receptor Splice Variant Is Up-Regulated during Prostate Cancer Progression and Promotes Androgen Depletion–Resistant Growth. Cancer Res. 2009, 69, 2305–2313. [Google Scholar] [CrossRef]
  22. Zhang, T.; Karsh, L.I.; Nissenblatt, M.J.; Canfield, S.E. Androgen Receptor Splice Variant, AR-V7, as a Biomarker of Resistance to Androgen Axis-Targeted Therapies in Advanced Prostate Cancer. Clin. Genitourin. Cancer 2020, 18, 1–10. [Google Scholar] [CrossRef] [PubMed]
  23. Saraon, P.; Drabovich, A.P.; Jarvi, K.A.; Diamandis, E.P. Mechanisms of Androgen-Independent Prostate Cancer. EJIFCC 2014, 25, 42–54. [Google Scholar] [PubMed]
  24. Heinlein, C.A.; Chang, C. Androgen Receptor (AR) Coregulators: An Overview. Endocr. Rev. 2002, 23, 175–200. [Google Scholar] [CrossRef] [PubMed]
  25. Gregory, C.W.; He, B.; Johnson, R.T.; Ford, O.H.; Mohler, J.L.; French, F.S.; Wilson, E.M. A Mechanism for Androgen Receptor-Mediated Prostate Cancer Recurrence after Androgen Deprivation Therapy. Cancer Res. 2001, 61, 4315–4319. [Google Scholar] [PubMed]
  26. Zhou, X.E.; Suino-Powell, K.M.; Li, J.; He, Y.; Mackeigan, J.P.; Melcher, K.; Yong, E.-L.; Xu, H.E. Identification of SRC3/AIB1 as a Preferred Coactivator for Hormone-Activated Androgen Receptor. J. Biol. Chem. 2010, 285, 9161–9171. [Google Scholar] [CrossRef] [PubMed]
  27. Tien, J.C.-Y.; Liu, Z.; Liao, L.; Wang, F.; Xu, Y.; Wu, Y.-L.; Zhou, N.; Ittmann, M.; Xu, J. The Steroid Receptor Coactivator-3 Is Required for the Development of Castration-Resistant Prostate Cancer. Cancer Res. 2013, 73, 3997–4008. [Google Scholar] [CrossRef] [PubMed]
  28. Cheng, S.; Brzostek, S.; Lee, S.R.; Hollenberg, A.N.; Balk, S.P. Inhibition of the Dihydrotestosterone-Activated Androgen Receptor by Nuclear Receptor Corepressor. Mol. Endocrinol. 2002, 16, 1492–1501. [Google Scholar] [CrossRef] [PubMed]
  29. Godoy, A.S.; Sotomayor, P.C.; Villagran, M.; Yacoub, R.; Montecinos, V.P.; McNerney, E.M.; Moser, M.; Foster, B.A.; Onate, S.A. Altered Corepressor SMRT Expression and Recruitment to Target Genes as a Mechanism That Change the Response to Androgens in Prostate Cancer Progression. Biochem. Biophys. Res. Commun. 2012, 423, 564–570. [Google Scholar] [CrossRef]
  30. Liao, G.; Chen, L.-Y.; Zhang, A.; Godavarthy, A.; Xia, F.; Ghosh, J.C.; Li, H.; Chen, J.D. Regulation of Androgen Receptor Activity by the Nuclear Receptor Corepressor SMRT*. J. Biol. Chem. 2003, 278, 5052–5061. [Google Scholar] [CrossRef]
  31. Nacusi, L.P.; Tindall, D.J. Targeting 5α-Reductase for Prostate Cancer Prevention and Treatment. Nat. Rev. Urol. 2011, 8, 378–384. [Google Scholar] [CrossRef]
  32. Dutt, S.S.; Gao, A.C. Molecular Mechanisms of Castration-Resistant Prostate Cancer Progression. Future Oncol. 2009, 5, 1403–1413. [Google Scholar] [CrossRef] [PubMed]
  33. Montgomery, R.B.; Mostaghel, E.A.; Vessella, R.; Hess, D.L.; Kalhorn, T.F.; Higano, C.S.; True, L.D.; Nelson, P.S. Maintenance of Intratumoral Androgens in Metastatic Prostate Cancer: A Mechanism for Castration-Resistant Tumor Growth. Cancer Res. 2008, 68, 4447–4454. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, J.D.; Haugk, K.; Woodke, L.; Nelson, P.; Coleman, I.; Plymate, S.R. Interaction of IGF Signaling and the Androgen Receptor in Prostate Cancer Progression. J. Cell. Biochem. 2006, 99, 392–401. [Google Scholar] [CrossRef] [PubMed]
  35. Malinowska, K.; Neuwirt, H.; Cavarretta, I.T.; Bektic, J.; Steiner, H.; Dietrich, H.; Moser, P.L.; Fuchs, D.; Hobisch, A.; Culig, Z. Interleukin-6 Stimulation of Growth of Prostate Cancer in Vitro and in Vivo through Activation of the Androgen Receptor. Endocr. Relat. Cancer 2009, 16, 155–169. [Google Scholar] [CrossRef] [PubMed]
  36. Sarker, D.; Reid, A.H.M.; Yap, T.A.; de Bono, J.S. Targeting the PI3K/AKT Pathway for the Treatment of Prostate Cancer. Clin. Cancer Res. 2009, 15, 4799–4805. [Google Scholar] [CrossRef] [PubMed]
  37. Wen, Y.; Hu, M.C.; Makino, K.; Spohn, B.; Bartholomeusz, G.; Yan, D.H.; Hung, M.C. HER-2/Neu Promotes Androgen-Independent Survival and Growth of Prostate Cancer Cells through the Akt Pathway. Cancer Res. 2000, 60, 6841–6845. [Google Scholar] [PubMed]
  38. Graff, J.R.; Konicek, B.W.; McNulty, A.M.; Wang, Z.; Houck, K.; Allen, S.; Paul, J.D.; Hbaiu, A.; Goode, R.G.; Sandusky, G.E.; et al. Increased AKT Activity Contributes to Prostate Cancer Progression by Dramatically Accelerating Prostate Tumor Growth and Diminishing p27Kip1 Expression. J. Biol. Chem. 2000, 275, 24500–24505. [Google Scholar] [CrossRef] [PubMed]
  39. Jamaspishvili, T.; Berman, D.M.; Ross, A.E.; Scher, H.I.; De Marzo, A.M.; Squire, J.A.; Lotan, T.L. Clinical Implications of PTEN Loss in Prostate Cancer. Nat. Rev. Urol. 2018, 15, 222–234. [Google Scholar] [CrossRef]
  40. Andrieu, C.; Taieb, D.; Baylot, V.; Ettinger, S.; Soubeyran, P.; De-Thonel, A.; Nelson, C.; Garrido, C.; So, A.; Fazli, L.; et al. Heat Shock Protein 27 Confers Resistance to Androgen Ablation and Chemotherapy in Prostate Cancer Cells through eIF4E. Oncogene 2010, 29, 1883–1896. [Google Scholar] [CrossRef]
  41. Baylot, V.; Katsogiannou, M.; Andrieu, C.; Taieb, D.; Acunzo, J.; Giusiano, S.; Fazli, L.; Gleave, M.; Garrido, C.; Rocchi, P. Targeting TCTP as a New Therapeutic Strategy in Castration-Resistant Prostate Cancer. Mol. Ther. J. Am. Soc. Gene Ther. 2012, 20, 2244–2256. [Google Scholar] [CrossRef]
  42. Cherif, C.; Nguyen, D.T.; Paris, C.; Le, T.K.; Sefiane, T.; Carbuccia, N.; Finetti, P.; Chaffanet, M.; Kaoutari, A.E.; Vernerey, J.; et al. Menin Inhibition Suppresses Castration-Resistant Prostate Cancer and Enhances Chemosensitivity. Oncogene 2022, 41, 125–137. [Google Scholar] [CrossRef] [PubMed]
  43. Le, T.K.; Cherif, C.; Omabe, K.; Paris, C.; Lannes, F.; Audebert, S.; Baudelet, E.; Hamimed, M.; Barbolosi, D.; Finetti, P.; et al. DDX5 mRNA-Targeting Antisense Oligonucleotide as a New Promising Therapeutic in Combating Castration-Resistant Prostate Cancer. Mol. Ther. 2022, 31, 471–486. [Google Scholar] [CrossRef] [PubMed]
  44. Arora, V.K.; Schenkein, E.; Murali, R.; Subudhi, S.K.; Wongvipat, J.; Balbas, M.D.; Shah, N.; Cai, L.; Efstathiou, E.; Logothetis, C.; et al. Glucocorticoid Receptor Confers Resistance to Antiandrogens by Bypassing Androgen Receptor Blockade. Cell 2013, 155, 1309–1322. [Google Scholar] [CrossRef] [PubMed]
  45. Puhr, M.; Hoefer, J.; Eigentler, A.; Ploner, C.; Handle, F.; Schaefer, G.; Kroon, J.; Leo, A.; Heidegger, I.; Eder, I.; et al. The Glucocorticoid Receptor Is a Key Player for Prostate Cancer Cell Survival and a Target for Improved Antiandrogen Therapy. Clin. Cancer Res. 2018, 24, 927–938. [Google Scholar] [CrossRef] [PubMed]
  46. Xie, N.; Cheng, H.; Lin, D.; Liu, L.; Yang, O.; Jia, L.; Fazli, L.; Gleave, M.E.; Wang, Y.; Rennie, P.; et al. The Expression of Glucocorticoid Receptor Is Negatively Regulated by Active Androgen Receptor Signaling in Prostate Tumors. Int. J. Cancer 2015, 136, E27–E38. [Google Scholar] [CrossRef] [PubMed]
  47. Patel, G.K.; Chugh, N.; Tripathi, M. Neuroendocrine Differentiation of Prostate Cancer—An Intriguing Example of Tumor Evolution at Play. Cancers 2019, 11, 1405. [Google Scholar] [CrossRef] [PubMed]
  48. Beltran, H.; Tomlins, S.; Aparicio, A.; Arora, V.; Rickman, D.; Ayala, G.; Huang, J.; True, L.; Gleave, M.E.; Soule, H.; et al. Aggressive Variants of Castration Resistant Prostate Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2014, 20, 2846–2850. [Google Scholar] [CrossRef] [PubMed]
  49. Beltran, H.; Rickman, D.S.; Park, K.; Chae, S.S.; Sboner, A.; MacDonald, T.Y.; Wang, Y.; Sheikh, K.L.; Terry, S.; Tagawa, S.T.; et al. Molecular Characterization of Neuroendocrine Prostate Cancer and Identification of New Drug Targets. Cancer Discov. 2011, 1, 487–495. [Google Scholar] [CrossRef]
  50. Lin, D.; Wyatt, A.W.; Xue, H.; Wang, Y.; Dong, X.; Haegert, A.; Wu, R.; Brahmbhatt, S.; Mo, F.; Jong, L.; et al. High Fidelity Patient-Derived Xenografts for Accelerating Prostate Cancer Discovery and Drug Development. Cancer Res. 2014, 74, 1272–1283. [Google Scholar] [CrossRef]
  51. Ku, S.Y.; Rosario, S.; Wang, Y.; Mu, P.; Seshadri, M.; Goodrich, Z.W.; Goodrich, M.M.; Labbé, D.P.; Gomez, E.C.; Wang, J.; et al. Rb1 and Trp53 Cooperate to Suppress Prostate Cancer Lineage Plasticity, Metastasis, and Antiandrogen Resistance. Science 2017, 355, 78–83. [Google Scholar] [CrossRef]
  52. Zou, M.; Toivanen, R.; Mitrofanova, A.; Floch, N.; Hayati, S.; Sun, Y.; Le Magnen, C.; Chester, D.; Mostaghel, E.A.; Califano, A.; et al. Transdifferentiation as a Mechanism of Treatment Resistance in a Mouse Model of Castration-Resistant Prostate Cancer. Cancer Discov. 2017, 7, 736–749. [Google Scholar] [CrossRef]
  53. Deng, X.; Liu, H.; Huang, J.; Cheng, L.; Keller, E.T.; Parsons, S.J.; Hu, C.-D. Ionizing Radiation Induces Prostate Cancer Neuroendocrine Differentiation through Interplay of CREB and ATF2: Implications for Disease Progression. Cancer Res. 2008, 68, 9663–9670. [Google Scholar] [CrossRef] [PubMed]
  54. Bonkhoff, H. Factors Implicated in Radiation Therapy Failure and Radiosensitization of Prostate Cancer. Prostate Cancer 2012, 2012, 593241. [Google Scholar] [CrossRef] [PubMed]
  55. Berruti, A.; Dogliotti, L.; Mosca, A.; Bellina, M.; Mari, M.; Torta, M.; Tarabuzzi, R.; Bollito, E.; Fontana, D.; Angeli, A. Circulating Neuroendocrine Markers in Patients with Prostate Carcinoma. Cancer 2000, 88, 2590–2597. [Google Scholar] [CrossRef] [PubMed]
  56. Davies, A.; Zoubeidi, A.; Selth, L.A. The Epigenetic and Transcriptional Landscape of Neuroendocrine Prostate Cancer. Endocr. Relat. Cancer 2020, 27, R35–R50. [Google Scholar] [CrossRef] [PubMed]
  57. Baca, S.C.; Takeda, D.Y.; Seo, J.-H.; Hwang, J.; Ku, S.Y.; Arafeh, R.; Arnoff, T.; Agarwal, S.; Bell, C.; O’Connor, E.; et al. Reprogramming of the FOXA1 Cistrome in Treatment-Emergent Neuroendocrine Prostate Cancer. Nat. Commun. 2021, 12, 1979. [Google Scholar] [CrossRef]
  58. Dang, Q.; Li, L.; Xie, H.; He, D.; Chen, J.; Song, W.; Chang, L.S.; Chang, H.-C.; Yeh, S.; Chang, C. Anti-Androgen Enzalutamide Enhances Prostate Cancer Neuroendocrine (NE) Differentiation via Altering the Infiltrated Mast Cells→Androgen Receptor (AR)→miRNA32 Signals. Mol. Oncol. 2015, 9, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  59. Mishra, R.; Haldar, S.; Placencio, V.; Madhav, A.; Rohena-Rivera, K.; Agarwal, P.; Duong, F.; Angara, B.; Tripathi, M.; Liu, Z.; et al. Stromal Epigenetic Alterations Drive Metabolic and Neuroendocrine Prostate Cancer Reprogramming. J. Clin. Investig. 2018, 128, 4472–4484. [Google Scholar] [CrossRef]
  60. Denmeade, S.R.; Isaacs, J.T. A History of Prostate Cancer Treatment. Nat. Rev. Cancer 2002, 2, 389–396. [Google Scholar] [CrossRef]
  61. US Food and Drug Administration. FDA Approves Relugolix for Advanced Prostate Cancer; US Food and Drug Administration: Silver Spring, MD, USA, 2021.
  62. Shore, N.D.; Saad, F.; Cookson, M.S.; George, D.J.; Saltzstein, D.R.; Tutrone, R.; Akaza, H.; Bossi, A.; van Veenhuyzen, D.F.; Selby, B.; et al. Oral Relugolix for Androgen-Deprivation Therapy in Advanced Prostate Cancer. N. Engl. J. Med. 2020, 382, 2187–2196. [Google Scholar] [CrossRef]
  63. Gamat, M.; McNeel, D.G. Androgen Deprivation and Immunotherapy for the Treatment of Prostate Cancer. Endocr. Relat. Cancer 2017, 24, T297–T310. [Google Scholar] [CrossRef] [PubMed]
  64. Merseburger, A.S.; Hammerer, P.; Rozet, F.; Roumeguère, T.; Caffo, O.; da Silva, F.C.; Alcaraz, A. Androgen Deprivation Therapy in Castrate-Resistant Prostate Cancer: How Important Is GnRH Agonist Backbone Therapy? World J. Urol. 2015, 33, 1079–1085. [Google Scholar] [CrossRef]
  65. Heidenreich, A.; Bastian, P.J.; Bellmunt, J.; Bolla, M.; Joniau, S.; van der Kwast, T.; Mason, M.; Matveev, V.; Wiegel, T.; Zattoni, F.; et al. EAU Guidelines on Prostate Cancer. Part II: Treatment of Advanced, Relapsing, and Castration-Resistant Prostate Cancer. Eur. Urol. 2014, 65, 467–479. [Google Scholar] [CrossRef] [PubMed]
  66. Cookson, M.S.; Roth, B.J.; Dahm, P.; Engstrom, C.; Freedland, S.J.; Hussain, M.; Lin, D.W.; Lowrance, W.T.; Murad, M.H.; Oh, W.K.; et al. Castration-Resistant Prostate Cancer: AUA Guideline. J. Urol. 2013, 190, 429–438. [Google Scholar] [CrossRef] [PubMed]
  67. Mohler, J.L.; Kantoff, P.W.; Armstrong, A.J.; Bahnson, R.R.; Cohen, M.; D’Amico, A.V.; Eastham, J.A.; Enke, C.A.; Farrington, T.A.; Higano, C.S.; et al. Prostate Cancer, Version 1.2014. J. Natl. Compr. Cancer Netw. 2013, 11, 1471–1479. [Google Scholar] [CrossRef] [PubMed]
  68. Crawford, E.D.; Rove, K.O. Incomplete Testosterone Suppression in Prostate Cancer. N. Engl. J. Med. 2010, 363, 1976. [Google Scholar] [CrossRef] [PubMed]
  69. Mostaghel, E.A.; Page, S.T.; Lin, D.W.; Fazli, L.; Coleman, I.M.; True, L.D.; Knudsen, B.; Hess, D.L.; Nelson, C.C.; Matsumoto, A.M.; et al. Intraprostatic Androgens and Androgen-Regulated Gene Expression Persist after Testosterone Suppression: Therapeutic Implications for Castration-Resistant Prostate Cancer. Cancer Res. 2007, 67, 5033–5041. [Google Scholar] [CrossRef]
  70. Egan, A.; Dong, Y.; Zhang, H.; Qi, Y.; Balk, S.P.; Sartor, O. Castration-Resistant Prostate Cancer: Adaptive Responses in the Androgen Axis. Cancer Treat. Rev. 2014, 40, 426–433. [Google Scholar] [CrossRef]
  71. Hager, S.; Ackermann, C.J.; Joerger, M.; Gillessen, S.; Omlin, A. Anti-Tumour Activity of Platinum Compounds in Advanced Prostate Cancer—A Systematic Literature Review. Ann. Oncol. 2016, 27, 975–984. [Google Scholar] [CrossRef]
  72. Martin, S.K.; Kyprianou, N. Chapter Three—Exploitation of the Androgen Receptor to Overcome Taxane Resistance in Advanced Prostate Cancer. In Advances in Cancer Research; Fisher, P.B., Tew, K.D., Eds.; Academic Press: Cambridge, MA, USA, 2015; Volume 127, pp. 123–158. [Google Scholar]
  73. Seruga, B.; Tannock, I.F. Chemotherapy-Based Treatment for Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2011, 29, 3686–3694. [Google Scholar] [CrossRef]
  74. Pivot, X.; Koralewski, P.; Hidalgo, J.L.; Chan, A.; Gonçalves, A.; Schwartsmann, G.; Assadourian, S.; Lotz, J.P. A Multicenter Phase II Study of XRP6258 Administered as a 1-h i.v. Infusion Every 3 Weeks in Taxane-Resistant Metastatic Breast Cancer Patients. Ann. Oncol. 2008, 19, 1547–1552. [Google Scholar] [CrossRef] [PubMed]
  75. Tannock, I.F.; Osoba, D.; Stockler, M.R.; Ernst, D.S.; Neville, A.J.; Moore, M.J.; Armitage, G.R.; Wilson, J.J.; Venner, P.M.; Coppin, C.M.; et al. Chemotherapy with Mitoxantrone plus Prednisone or Prednisone Alone for Symptomatic Hormone-Resistant Prostate Cancer: A Canadian Randomized Trial with Palliative End Points. J. Clin. Oncol. 1996, 14, 1756–1764. Available online: https://ascopubs.org/doi/10.1200/JCO.1996.14.6.1756 (accessed on 25 May 2023). [CrossRef] [PubMed]
  76. Mitoxantrone. Available online: https://go.drugbank.com/drugs/DB01204 (accessed on 24 May 2023).
  77. Chemotherapy for Prostate Cancer. Available online: https://www.cancer.org/cancer/types/prostate-cancer/treating/chemotherapy.html (accessed on 15 August 2023).
  78. Rehman, L.U.; Nisar, M.H.; Fatima, W.; Sarfraz, A.; Azeem, N.; Sarfraz, Z.; Robles-Velasco, K.; Cherrez-Ojeda, I. Immunotherapy for Prostate Cancer: A Current Systematic Review and Patient Centric Perspectives. J. Clin. Med. 2023, 12, 1446. [Google Scholar] [CrossRef] [PubMed]
  79. Westdorp, H.; Sköld, A.E.; Snijer, B.A.; Franik, S.; Mulder, S.F.; Major, P.P.; Foley, R.; Gerritsen, W.R.; de Vries, I.J.M. Immunotherapy for Prostate Cancer: Lessons from Responses to Tumor-Associated Antigens. Front. Immunol. 2014, 5, 191. [Google Scholar] [CrossRef] [PubMed]
  80. Dendreon Corporation. Sipuleucel-T. Drugs R D 2006, 7, 197–201. [Google Scholar] [CrossRef] [PubMed]
  81. Fay, E.K.; Graff, J.N. Immunotherapy in Prostate Cancer. Cancers 2020, 12, 1752. [Google Scholar] [CrossRef] [PubMed]
  82. Buonerba, C.; Ferro, M.; Di Lorenzo, G. Sipuleucel-T for Prostate Cancer: The Immunotherapy Era Has Commenced. Expert Rev. Anticancer Ther. 2011, 11, 25–28. [Google Scholar] [CrossRef] [PubMed]
  83. Kantoff, P.W.; Higano, C.S.; Shore, N.D.; Berger, E.R.; Small, E.J.; Penson, D.F.; Redfern, C.H.; Ferrari, A.C.; Dreicer, R.; Sims, R.B.; et al. Sipuleucel-T Immunotherapy for Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2010, 363, 411–422. [Google Scholar] [CrossRef]
  84. Iranzo, P.; Callejo, A.; Assaf, J.D.; Molina, G.; Lopez, D.E.; Garcia-Illescas, D.; Pardo, N.; Navarro, A.; Martinez-Marti, A.; Cedres, S.; et al. Overview of Checkpoint Inhibitors Mechanism of Action: Role of Immune-Related Adverse Events and Their Treatment on Progression of Underlying Cancer. Front. Med. 2022, 9, 875974. [Google Scholar] [CrossRef]
  85. Marcus, L.; Lemery, S.J.; Keegan, P.; Pazdur, R. FDA Approval Summary: Pembrolizumab for the Treatment of Microsatellite Instability-High Solid Tumors. Clin. Cancer Res. 2019, 25, 3753–3758. [Google Scholar] [CrossRef]
  86. Shimizu, K.; Sano, T.; Mizuno, K.; Sunada, T.; Makita, N.; Hagimoto, H.; Goto, T.; Sawada, A.; Fujimoto, M.; Ichioka, K.; et al. A Case of Microsatellite Instability-High Clinically Advanced Castration-Resistant Prostate Cancer Showing a Remarkable Response to Pembrolizumab Sustained over at Least 18 Months. Cold Spring Harb. Mol. Case Stud. 2022, 8, a006194. [Google Scholar] [CrossRef] [PubMed]
  87. Graff, J.N.; Beer, T.M.; Alumkal, J.J.; Slottke, R.E.; Redmond, W.L.; Thomas, G.V.; Thompson, R.F.; Wood, M.A.; Koguchi, Y.; Chen, Y.; et al. A Phase II Single-Arm Study of Pembrolizumab with Enzalutamide in Men with Metastatic Castration-Resistant Prostate Cancer Progressing on Enzalutamide Alone. J. Immunother. Cancer 2020, 8, e000642. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, I.; Song, L.; Wang, B.Y.; Kalebasty, A.R.; Uchio, E.; Zi, X. Prostate Cancer Immunotherapy: A Review of Recent Advancements with Novel Treatment Methods and Efficacy. Am. J. Clin. Exp. Urol. 2022, 10, 210–233. [Google Scholar] [PubMed]
  89. Israeli, R.S.; Powell, C.T.; Corr, J.G.; Fair, W.R.; Heston, W.D. Expression of the Prostate-Specific Membrane Antigen. Cancer Res. 1994, 54, 1807–1811. [Google Scholar]
  90. Palamiuc, L.; Emerling, B.M. PSMA Brings New Flavors to PI3K Signaling: A Role for Glutamate in Prostate Cancer. J. Exp. Med. 2018, 215, 17–19. [Google Scholar] [CrossRef] [PubMed]
  91. Kaittanis, C.; Andreou, C.; Hieronymus, H.; Mao, N.; Foss, C.A.; Eiber, M.; Weirich, G.; Panchal, P.; Gopalan, A.; Zurita, J.; et al. Prostate-Specific Membrane Antigen Cleavage of Vitamin B9 Stimulates Oncogenic Signaling through Metabotropic Glutamate Receptors. J. Exp. Med. 2017, 215, 159–175. [Google Scholar] [CrossRef] [PubMed]
  92. Weineisen, M.; Schottelius, M.; Simecek, J.; Baum, R.P.; Yildiz, A.; Beykan, S.; Kulkarni, H.R.; Lassmann, M.; Klette, I.; Eiber, M.; et al. 68Ga- and 177Lu-Labeled PSMA I&T: Optimization of a PSMA-Targeted Theranostic Concept and First Proof-of-Concept Human Studies. J. Nucl. Med. 2015, 56, 1169–1176. [Google Scholar] [CrossRef]
  93. Hofman, M.S.; Emmett, L.; Sandhu, S.; Iravani, A.; Joshua, A.M.; Goh, J.C.; Pattison, D.A.; Tan, T.H.; Kirkwood, I.D.; Ng, S.; et al. [177Lu]Lu-PSMA-617 versus Cabazitaxel in Patients with Metastatic Castration-Resistant Prostate Cancer (TheraP): A Randomised, Open-Label, Phase 2 Trial. Lancet 2021, 397, 797–804. [Google Scholar] [CrossRef] [PubMed]
  94. Fallah, J.; Agrawal, S.; Gittleman, H.; Fiero, M.H.; Subramaniam, S.; John, C.; Chen, W.; Ricks, T.K.; Niu, G.; Fotenos, A.; et al. FDA Approval Summary: Lutetium Lu 177 Vipivotide Tetraxetan for Patients with Metastatic Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2023, 29, 1651–1657. [Google Scholar] [CrossRef]
  95. EMA Pluvicto. Available online: https://www.ema.europa.eu/en/medicines/human/EPAR/pluvicto (accessed on 15 May 2023).
  96. Sartor, A.O.; Morris, M.J.; Messman, R.; Krause, B.J. VISION: An International, Prospective, Open-Label, Multicenter, Randomized Phase III Study of 177Lu-PSMA-617 in the Treatment of Patients with Progressive PSMA-Positive Metastatic Castration-Resistant Prostate Cancer (mCRPC). JCO 2020, 38, TPS259. [Google Scholar] [CrossRef]
  97. Sartor, O.; de Bono, J.; Chi, K.N.; Fizazi, K.; Herrmann, K.; Rahbar, K.; Tagawa, S.T.; Nordquist, L.T.; Vaishampayan, N.; El-Haddad, G.; et al. Lutetium-177–PSMA-617 for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2021, 385, 1091–1103. [Google Scholar] [CrossRef] [PubMed]
  98. Gafita, A.; Marcus, C.; Kostos, L.; Schuster, D.M.; Calais, J.; Hofman, M.S. Predictors and Real-World Use of Prostate-Specific Radioligand Therapy: PSMA and Beyond. Am. Soc. Clin. Oncol. Educ. Book 2022, 42, 366–382. [Google Scholar] [CrossRef] [PubMed]
  99. Baboudjian, M.; Gauthé, M.; Barret, E.; Brureau, L.; Rocchi, P.; Créhange, G.; Dariane, C.; Fiard, G.; Fromont, G.; Beauval, J.-B.; et al. How PET-CT Is Changing the Management of Non-Metastatic Castration-Resistant Prostate Cancer?: Comment La TEP-TDM Peut Modifier La Prise En Charge Du Cancer de La Prostate Non Métastatique Résistant à La Castration? Prog. Urol. 2022, 32, 6S43–6S53. [Google Scholar] [CrossRef] [PubMed]
  100. Gafita, A.; Rauscher, I.; Weber, M.; Hadaschik, B.; Wang, H.; Armstrong, W.R.; Tauber, R.; Grogan, T.R.; Czernin, J.; Rettig, M.B.; et al. Novel Framework for Treatment Response Evaluation Using PSMA PET/CT in Patients with Metastatic Castration-Resistant Prostate Cancer (RECIP 1.0): An International Multicenter Study. J. Nucl. Med. 2022, 63, 1651–1658. [Google Scholar] [CrossRef] [PubMed]
  101. Buteau, J.P.; Martin, A.J.; Emmett, L.; Iravani, A.; Sandhu, S.; Joshua, A.M.; Francis, R.J.; Zhang, A.Y.; Scott, A.M.; Lee, S.-T.; et al. PSMA and FDG-PET as Predictive and Prognostic Biomarkers in Patients given [177Lu]Lu-PSMA-617 versus Cabazitaxel for Metastatic Castration-Resistant Prostate Cancer (TheraP): A Biomarker Analysis from a Randomised, Open-Label, Phase 2 Trial. Lancet Oncol. 2022, 23, 1389–1397. [Google Scholar] [CrossRef] [PubMed]
  102. John, N.; Pathmanandavel, S.; Crumbaker, M.; Counter, W.; Ho, B.; Yam, A.O.; Wilson, P.; Niman, R.; Ayers, M.; Poole, A.; et al. 177Lu-PSMA SPECT Quantitation at 6 Weeks (Dose 2) Predicts Short Progression Free Survival for Patients Undergoing Lu PSMA I&T Therapy. J. Nucl. Med. 2022, 64, 610–615. [Google Scholar] [CrossRef]
  103. Rosado, M.M.; Pioli, C. Radiotherapy, PARP Inhibition, and Immune-Checkpoint Blockade: A Triad to Overcome the Double-Edged Effects of Each Single Player. Cancers 2023, 15, 1093. [Google Scholar] [CrossRef]
  104. Khreish, F.; Ebert, N.; Ries, M.; Maus, S.; Rosar, F.; Bohnenberger, H.; Stemler, T.; Saar, M.; Bartholomä, M.; Ezziddin, S. 225Ac-PSMA-617/177Lu-PSMA-617 Tandem Therapy of Metastatic Castration-Resistant Prostate Cancer: Pilot Experience. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 721–728. [Google Scholar] [CrossRef]
  105. Feuerecker, B.; Tauber, R.; Knorr, K.; Heck, M.; Beheshti, A.; Seidl, C.; Bruchertseifer, F.; Pickhard, A.; Gafita, A.; Kratochwil, C.; et al. Activity and Adverse Events of Actinium-225-PSMA-617 in Advanced Metastatic Castration-Resistant Prostate Cancer After Failure of Lutetium-177-PSMA. Eur. Urol. 2021, 79, 343–350. [Google Scholar] [CrossRef]
  106. Paschalis, A.; Sheehan, B.; Riisnaes, R.; Rodrigues, D.N.; Gurel, B.; Bertan, C.; Ferreira, A.; Lambros, M.B.K.; Seed, G.; Yuan, W.; et al. Prostate-Specific Membrane Antigen Heterogeneity and DNA Repair Defects in Prostate Cancer. Eur. Urol. 2019, 76, 469–478. [Google Scholar] [CrossRef]
  107. Bakht, M.K.; Yamada, Y.; Ku, S.-Y.; Venkadakrishnan, V.B.; Korsen, J.A.; Kalidindi, T.M.; Mizuno, K.; Ahn, S.H.; Seo, J.-H.; Garcia, M.M.; et al. Landscape of Prostate-Specific Membrane Antigen Heterogeneity and Regulation in AR-Positive and AR-Negative Metastatic Prostate Cancer. Nat. Cancer 2023, 4, 699–715. [Google Scholar] [CrossRef] [PubMed]
  108. Taïeb, D.; Foletti, J.-M.; Bardiès, M.; Rocchi, P.; Hicks, R.J.; Haberkorn, U. PSMA-Targeted Radionuclide Therapy and Salivary Gland Toxicity: Why Does It Matter? J. Nucl. Med. 2018, 59, 747–748. [Google Scholar] [CrossRef] [PubMed]
  109. Gupta, N.; Devgan, A.; Bansal, I.; Olsavsky, T.D.; Li, S.; Abdelbaki, A.; Kumar, Y. Usefulness of Radium-223 in Patients with Bone Metastases. Proc. Bayl. Univ. Med. Cent. 2017, 30, 424–426. [Google Scholar] [CrossRef] [PubMed]
  110. Ritter, M.A.; Cleaver, J.E.; Tobias, C.A. High-LET Radiations Induce a Large Proportion of Non-Rejoining DNA Breaks. Nature 1977, 266, 653–655. [Google Scholar] [CrossRef] [PubMed]
  111. Parker, C.; Nilsson, S.; Heinrich, D.; Helle, S.I.; O’Sullivan, J.M.; Fosså, S.D.; Chodacki, A.; Wiechno, P.; Logue, J.; Seke, M.; et al. Alpha Emitter Radium-223 and Survival in Metastatic Prostate Cancer. N. Engl. J. Med. 2013, 369, 213–223. [Google Scholar] [CrossRef] [PubMed]
  112. Nilsson, S.; Cislo, P.; Sartor, O.; Vogelzang, N.J.; Coleman, R.E.; O’Sullivan, J.M.; Reuning-Scherer, J.; Shan, M.; Zhan, L.; Parker, C. Patient-Reported Quality-of-Life Analysis of Radium-223 Dichloride from the Phase III ALSYMPCA Study. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2016, 27, 868–874. [Google Scholar] [CrossRef]
  113. Congregado, B.; Rivero, I.; Osmán, I.; Sáez, C.; López, R.M. PARP Inhibitors: A New Horizon for Patients with Prostate Cancer. Biomedicines 2022, 10, 1416. [Google Scholar] [CrossRef]
  114. Liang, F.; Han, M.; Romanienko, P.J.; Jasin, M. Homology-Directed Repair Is a Major Double-Strand Break Repair Pathway in Mammalian Cells. Proc. Natl. Acad. Sci. USA 1998, 95, 5172–5177. [Google Scholar] [CrossRef]
  115. FDA Approves Olaparib, Rucaparib to Treat Prostate Cancer—NCI. Available online: https://www.cancer.gov/news-events/cancer-currents-blog/2020/fda-olaparib-rucaparib-prostate-cancer (accessed on 12 May 2023).
  116. De Bono, J.; Mateo, J.; Fizazi, K.; Saad, F.; Shore, N.; Sandhu, S.; Chi, K.N.; Sartor, O.; Agarwal, N.; Olmos, D.; et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2020, 382, 2091–2102. [Google Scholar] [CrossRef]
  117. Fizazi, K.; Piulats, J.M.; Reaume, M.N.; Ostler, P.; McDermott, R.; Gingerich, J.R.; Pintus, E.; Sridhar, S.S.; Bambury, R.M.; Emmenegger, U.; et al. Rucaparib or Physician’s Choice in Metastatic Prostate Cancer. N. Engl. J. Med. 2023, 388, 719–732. [Google Scholar] [CrossRef]
  118. US Food and Drug Administration. FDA Approves Olaparib with Abiraterone and Prednisone (or Prednisolone) for BRCA-Mutated Metastatic Castration-Resistant Prostate Cancer; US Food and Drug Administration: Silver Spring, MD, USA, 2023.
  119. Smith, M.R.; Scher, H.I.; Sandhu, S.; Efstathiou, E.; Lara, P.N.; Yu, E.Y.; George, D.J.; Chi, K.N.; Saad, F.; Ståhl, O.; et al. Niraparib in Patients with Metastatic Castration-Resistant Prostate Cancer and DNA Repair Gene Defects (GALAHAD): A Multicentre, Open-Label, Phase 2 Trial. Lancet Oncol. 2022, 23, 362–373. [Google Scholar] [CrossRef]
  120. De Bono, J.S.; Mehra, N.; Scagliotti, G.V.; Castro, E.; Dorff, T.; Stirling, A.; Stenzl, A.; Fleming, M.T.; Higano, C.S.; Saad, F.; et al. Talazoparib Monotherapy in Metastatic Castration-Resistant Prostate Cancer with DNA Repair Alterations (TALAPRO-1): An Open-Label, Phase 2 Trial. Lancet Oncol. 2021, 22, 1250–1264. [Google Scholar] [CrossRef]
  121. Agarwal, N.; Azad, A.A.; Carles, J.; Fay, A.P.; Matsubara, N.; Heinrich, D.; Szczylik, C.; Giorgi, U.D.; Joung, J.Y.; Fong, P.C.C.; et al. Talazoparib plus Enzalutamide in Men with First-Line Metastatic Castration-Resistant Prostate Cancer (TALAPRO-2): A Randomised, Placebo-Controlled, Phase 3 Trial. Lancet 2023, 402, 291–303. [Google Scholar] [CrossRef]
  122. US Food and Drug Administration. FDA Approves Niraparib and Abiraterone Acetate plus Prednisone for BRCA-Mutated Metastatic Castration-Resistant Prostate Cancer; US Food and Drug Administration: Silver Spring, MD, USA, 2023.
  123. Slootbeek, P.H.J.; Overbeek, J.K.; Ligtenberg, M.J.L.; van Erp, N.P.; Mehra, N. PARPing up the Right Tree; an Overview of PARP Inhibitors for Metastatic Castration-Resistant Prostate Cancer. Cancer Lett. 2023, 216367. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of castration-resistant prostate cancer (CRPC). AR overexpression enables the survival and proliferation of tumor cells in limited-androgen conditions during androgen suppression treatment. A high AR expression level can be due to AR gene amplification, epigenetics, and miRNA modulation (1). Point mutations in the ligand-binding domain of the AR gene lead to the increased affinity of the mutated AR (mAR) to other hormones, such as progesterone and estrogen, thereby modulating androgen-responsive gene transcription independently from androgen (2). The emergence of AR-splicing isoforms, such as AR3 (also called AR-V7), AR4, and AR5, encoding the truncated AR protein (tAR), which lacks the ligand-binding domain, results in constitutive activation of the AR, thereby promoting variant-carrying tumor cells to ignore the need for androgen (3). Overexpression of AR coactivators and the decreased expression of AR corepressors will result in an increase in AR-regulated transcription (4). Increased production of 5α-reductase can provide sufficient androgens for AR activation in cancer cells (5). CRPC can be induced by outlaw pathways in which AR signaling can be activated in a ligand-independent manner by other molecules than androgens, such as growth factors, cytokines, and kinases (6). The bypass pathways, involved in CRPC progression, increase the activity of MAPK, PI3K, and PCK cascades, leading to either the stimulation of alternative growth pathways or the enhancement of survival signaling independently from AR signaling. Alternative pathways also involve the overexpression of heat-shock protein 27 (HSP27), which mediates its cytoprotective function by protecting its interacting proteins (eIF4E, TCTP, Menin, DDX5) from their degradation by the proteasome (7) (↑ in red: increase, ↓in red: decrease).
Figure 1. Mechanisms of castration-resistant prostate cancer (CRPC). AR overexpression enables the survival and proliferation of tumor cells in limited-androgen conditions during androgen suppression treatment. A high AR expression level can be due to AR gene amplification, epigenetics, and miRNA modulation (1). Point mutations in the ligand-binding domain of the AR gene lead to the increased affinity of the mutated AR (mAR) to other hormones, such as progesterone and estrogen, thereby modulating androgen-responsive gene transcription independently from androgen (2). The emergence of AR-splicing isoforms, such as AR3 (also called AR-V7), AR4, and AR5, encoding the truncated AR protein (tAR), which lacks the ligand-binding domain, results in constitutive activation of the AR, thereby promoting variant-carrying tumor cells to ignore the need for androgen (3). Overexpression of AR coactivators and the decreased expression of AR corepressors will result in an increase in AR-regulated transcription (4). Increased production of 5α-reductase can provide sufficient androgens for AR activation in cancer cells (5). CRPC can be induced by outlaw pathways in which AR signaling can be activated in a ligand-independent manner by other molecules than androgens, such as growth factors, cytokines, and kinases (6). The bypass pathways, involved in CRPC progression, increase the activity of MAPK, PI3K, and PCK cascades, leading to either the stimulation of alternative growth pathways or the enhancement of survival signaling independently from AR signaling. Alternative pathways also involve the overexpression of heat-shock protein 27 (HSP27), which mediates its cytoprotective function by protecting its interacting proteins (eIF4E, TCTP, Menin, DDX5) from their degradation by the proteasome (7) (↑ in red: increase, ↓in red: decrease).
Cancers 15 05047 g001
Figure 2. Current therapies for castration-resistant prostate cancer (CRPC) treatment. Along with androgen deprivation therapy (ADT), the approved therapies by the FDA and EMA to treat CRPC include chemotherapy (A), immunotherapy (B), radiotherapy (C), and PARP inhibitors (D). (A) Chemotherapy: (1) Docetaxel and cabazitaxel prevent microtube depolymerization, therefore inhibiting cell division and causing cell death. Mitoxantrone could induce cell death via the inhibition of topoisomerase II (2) and DNA damage induction (3). (B) Immunotherapy: Sipuleucel-T stimulates the T-cell anti-tumor activity by targeting the prostatic acid phosphatase (PAP) protein—an overexpressed protein in prostate cancer cells. (1) Antigen-presenting cells (APCs) isolated from patients will digest PAP proteins into small peptides and display these PAP peptides on their surfaces. (2) APCs present PAP peptides to T-cells, which can then recognize cancer cells that express PAP on their surfaces and activate immune cytotoxic effects to kill CRPC cells (3). Immune checkpoint inhibitors, including pembrolizumab and dostarlimab, are utilized to combat cancer cells by blocking immune checkpoint pathways, which reactivates cytotoxic lymphocytes antitumor responses. These medications work by preventing interaction between the PD-1 receptor and its ligands (PD-L1 and PD-L2), consequently inhibiting the programmed death 1 pathway and triggering the immune response against cancer cells. (C) Targeted Radiation Therapy: (1) Radium-223 induces cell death by generating DNA damage via emitting high-energy α-particles; (2) [177Lutetium]-PSMA-617 contains the ligand of the prostate-specific membrane antigen of prostate cancer cells (PSMA-617) conjugated with radiolabeled 177Lutetium. This drug causes DNA damage via the release of Ɓ and γ particles, thus leading to cell death. (D) PARP Inhibitors: The inhibition of PARP enzymes leads to double-strand DNA breaks from single-strand DNA breaks (SSBs). (1) The BRCA1/2 gene will be activated to repair DSBs and promote cell survival. (2) Cells with BRCA1/2 alterations cannot repair DSBs, leading to DNA damage and inducing cell death.
Figure 2. Current therapies for castration-resistant prostate cancer (CRPC) treatment. Along with androgen deprivation therapy (ADT), the approved therapies by the FDA and EMA to treat CRPC include chemotherapy (A), immunotherapy (B), radiotherapy (C), and PARP inhibitors (D). (A) Chemotherapy: (1) Docetaxel and cabazitaxel prevent microtube depolymerization, therefore inhibiting cell division and causing cell death. Mitoxantrone could induce cell death via the inhibition of topoisomerase II (2) and DNA damage induction (3). (B) Immunotherapy: Sipuleucel-T stimulates the T-cell anti-tumor activity by targeting the prostatic acid phosphatase (PAP) protein—an overexpressed protein in prostate cancer cells. (1) Antigen-presenting cells (APCs) isolated from patients will digest PAP proteins into small peptides and display these PAP peptides on their surfaces. (2) APCs present PAP peptides to T-cells, which can then recognize cancer cells that express PAP on their surfaces and activate immune cytotoxic effects to kill CRPC cells (3). Immune checkpoint inhibitors, including pembrolizumab and dostarlimab, are utilized to combat cancer cells by blocking immune checkpoint pathways, which reactivates cytotoxic lymphocytes antitumor responses. These medications work by preventing interaction between the PD-1 receptor and its ligands (PD-L1 and PD-L2), consequently inhibiting the programmed death 1 pathway and triggering the immune response against cancer cells. (C) Targeted Radiation Therapy: (1) Radium-223 induces cell death by generating DNA damage via emitting high-energy α-particles; (2) [177Lutetium]-PSMA-617 contains the ligand of the prostate-specific membrane antigen of prostate cancer cells (PSMA-617) conjugated with radiolabeled 177Lutetium. This drug causes DNA damage via the release of Ɓ and γ particles, thus leading to cell death. (D) PARP Inhibitors: The inhibition of PARP enzymes leads to double-strand DNA breaks from single-strand DNA breaks (SSBs). (1) The BRCA1/2 gene will be activated to repair DSBs and promote cell survival. (2) Cells with BRCA1/2 alterations cannot repair DSBs, leading to DNA damage and inducing cell death.
Cancers 15 05047 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Le, T.K.; Duong, Q.H.; Baylot, V.; Fargette, C.; Baboudjian, M.; Colleaux, L.; Taïeb, D.; Rocchi, P. Castration-Resistant Prostate Cancer: From Uncovered Resistance Mechanisms to Current Treatments. Cancers 2023, 15, 5047. https://doi.org/10.3390/cancers15205047

AMA Style

Le TK, Duong QH, Baylot V, Fargette C, Baboudjian M, Colleaux L, Taïeb D, Rocchi P. Castration-Resistant Prostate Cancer: From Uncovered Resistance Mechanisms to Current Treatments. Cancers. 2023; 15(20):5047. https://doi.org/10.3390/cancers15205047

Chicago/Turabian Style

Le, Thi Khanh, Quang Hieu Duong, Virginie Baylot, Christelle Fargette, Michael Baboudjian, Laurence Colleaux, David Taïeb, and Palma Rocchi. 2023. "Castration-Resistant Prostate Cancer: From Uncovered Resistance Mechanisms to Current Treatments" Cancers 15, no. 20: 5047. https://doi.org/10.3390/cancers15205047

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop